首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
This work reports the preparation of two recombinant strains each containing two enzymatic activities mutually expressed through regulated systems for production of functionalized epoxides in one-pot reactions. One strain was Pseudomonas putida PaW340, containing the gene coding for styrene monooxygenase (SMO) from Pseudomonas fluorescens ST under the auto-inducing Ptou promoter and the TouR regulator of Pseudomonas sp. OX1 and the gene coding for naphthalene dihydrodiol dehydrogenase (NDDH) from P. fluorescens N3 under the Ptac promoter inducible by IPTG. The second strain was Escherichia coli JM109, in which the expression of SMO was under the control of the Pnah promoter and the NahR regulator of P. fluorescens N3 inducible by salicylate, while the gene expressing NDDH was under the control of the Plac promoter inducible by IPTG. SMO and NDDH activities were tested in bioconversion experiments using cinnamyl alcohol as reference substrate. The application that we selected is one example of the sequential use of the two enzymatic activities which require a temporal control of the expression of both genes.  相似文献   

2.
(S)-Styrene oxide and (R)-1,2-phenylethanediol are chiral aromatic molecular building blocks used commonly as precursors to pharmaceuticals and other specialty chemicals. Two pathways have been engineered in Escherichia coli for their individual biosynthesis directly from glucose. The novel pathways each constitute extensions of the previously engineered styrene pathway, developed by co-expressing either styrene monooxygenase (SMO) or styrene dioxygenase (SDO) to convert styrene to (S)-styrene oxide and (R)-1,2-phenylethanediol, respectively. StyAB from Pseudomonas putida S12 was determined to be the most effective SMO. SDO activity was achieved using NahAaAbAcAd of Pseudomonas sp. NCIB 9816-4, a naphthalene dioxygenase with known broad substrate specificity. Production of phenylalanine, the precursor to both pathways, was systematically enhanced through a number of mutations, most notably via deletion of tyrA and over-expression of tktA. As a result, (R)-1,2-phenylethanediol reached titers as high as 1.23 g/L, and at 1.32 g/L (S)-styrene oxide titers already approach their toxicity limit. As with other aromatics, product toxicity was strongly correlated with a model of membrane accumulation and disruption. This study additionally demonstrates that greater flux through the styrene pathway can be achieved if its toxicity is addressed, as achieved in this case by reacting styrene to less toxic products. See accompanying commentary by Brian F. Pfleger DOI: 10.1002/biot.201300251  相似文献   

3.
4.
Recently isolated,Pseudomonas putida SN1 grows on styrene as its sole carbon and energy source through successive oxidation of styrene by styrene monooxygenase (SMO), styrene oxide isomerase (SOI), and phenylacetaldehyde dehydrogenase. For the production of (S)-styrene oxide, two knockout mutants of SN1 were constructed, one lacking SOI and another lacking both SMO and SOI. These mutants were developed into whole-cell biocatalysts by transformation with a multicopy plasmid vector containing SMO genes (styAB) of the SN1. Neither of these self-cloned recombinants could grow on styrene, but both converted styrene into an enantiopure (S)-styrene oxide (e.e.>99%). Whole-cell SMO activity was higher in the recombinant constructed from the SOI-deleted mutant (130 U/g cdw) than in the other one (35 U/g cdw). However, the SMO activity of the former was about the same as that of the SOI-deleted SN1 possessing a single copy of thestyAB gene that was used as host. This indicates that the copy number ofstyAB genes is not rate-limiting on SMO catalysis by whole-cell SN1.  相似文献   

5.
Sequence analysis of a 9-kb genomic fragment of the actinobacterium Rhodococcus opacus 1CP led to identification of an open reading frame encoding a novel fusion protein, StyA2B, with a putative function in styrene metabolism via styrene oxide and phenylacetic acid. Gene cluster analysis indicated that the highly related fusion proteins of Nocardia farcinica IFM10152 and Arthrobacter aurescens TC1 are involved in a similar physiological process. Whereas 413 amino acids of the N terminus of StyA2B are highly similar to those of the oxygenases of two-component styrene monooxygenases (SMOs) from pseudomonads, the residual 160 amino acids of the C terminus show significant homology to the flavin reductases of these systems. Cloning and functional expression of His10-StyA2B revealed for the first time that the fusion protein does in fact catalyze two separate reactions. Strictly NADH-dependent reduction of flavins and highly enantioselective oxygenation of styrene to (S)-styrene oxide were shown. Inhibition studies and photometric analysis of recombinant StyA2B indicated the absence of tightly bound heme and flavin cofactors in this self-sufficient monooxygenase. StyA2B oxygenates a spectrum of aromatic compounds similar to those of two-component SMOs. However, the specific activities of the flavin-reducing and styrene-oxidizing functions of StyA2B are one to two orders of magnitude lower than those of StyA/StyB from Pseudomonas sp. strain VLB120.The incorporation of one atom of oxygen during hydroxylation, epoxidation, sulfoxidation, or Baeyer-Villiger oxidation is a common initial step of the aerobic degradation of aromatic compounds by microorganisms. In bacteria, these reactions are most frequently catalyzed by inducible flavoprotein monooxygenases (EC 1.14.13 [57]). The majority of these enzymes (so-called single-component flavoprotein monooxygenases) utilize electrons from NAD(P)H, which are transferred to a non-covalently bound flavin adenine dinucleotide (FAD) in order to activate molecular oxygen as a flavin (hydro)peroxide. Depending on the protonation of this intermediate and the type of substrate, an oxygen atom is then incorporated by nucleophilic or electrophilic attack. More recently, different two-component flavoprotein monooxygenases have been characterized (57). These systems cover an NAD(P)H-dependent flavin reductase in order to generate reduced flavin and an oxygenase that utilizes this cofactor for the activation of oxygen.The exquisite regio- and stereoselectivities of oxygen insertion by flavoprotein monooxygenases favor these enzymes for biocatalytic applications (23, 24, 33). This is especially true because chemical synthesis approaches by hetero- or homogenic catalysis often do not yield a sufficiently high enantiomeric excess for the production of pharmaceuticals and their chiral building blocks. The use of oxygen as an inexpensive nontoxic oxidant and mild reaction conditions are additional advantages with the potential for increasing the environmental sustainability of oxygenase-catalyzed biotransformations.The necessity for expensive cofactors is perhaps the most striking drawback limiting the industrial application of flavoprotein monooxygenases. Different electrochemical and enzymatic procedures for in vitro cofactor regeneration are available (20, 21, 32, 52, 56), but these systems are currently lacking long-term stability. As a consequence, the practical application of flavoprotein monooxygenases is virtually restricted to in vivo systems in which cofactor regeneration is mediated by the metabolism of the expression host (45, 49). The limitations of whole-cell-mediated biotransformations by substrate and/or product toxicity can be overcome by means of two-phase systems, as was recently shown for the two-component styrene monooxygenase (SMO) from Pseudomonas sp. strain VLB120 (45).Two-component flavoprotein monooxygenases present additional challenges for biocatalytic applications. The need for two separate protein components may hamper attempts at recombinant enzyme expression, the application of immobilized enzymes in cell-free systems, and the detection of novel oxygenases during activity-based metagenome-screening approaches. Moreover, and as was already shown for the two-component SMO from Pseudomonas sp. strain VLB120 (44), the interprotein transfer of reduced FAD is accompanied to a certain extent by the auto-oxidative formation of reactive oxygen species such as hydrogen peroxide (Fig. (Fig.1).1). Auto-oxidation of reduced FAD not only decreases the efficiency of the oxygenation process but also negatively interferes with the physiological conditions of the cell. The extent of oxidative stress is considerably increased when FAD oxidoreductase activity exceeds oxygenase activity and uncoupling becomes dominant.Open in a separate windowFIG. 1.Catalytic mechanism of two-component SMOs and the formation of oxidative stress by uncoupling between FAD oxidoreductase (StyB) and oxygenase (StyA) (adapted from reference 36). FADOX and FADRED, oxidized and reduced forms of FAD, respectively.Presently the details of reduced-FAD transfer between the oxygenase and FAD oxidoreductase components of SMOs are not known. Recent kinetic studies have indicated that reduced FAD is transferred by a mixed mechanism in which direct contact of both proteins and free diffusion of the reduced cofactor play a role (25). This hypothesis is supported by the work of Otto and coworkers in which the formation of hydrogen peroxide was shown to be reciprocally proportional to the concentration of active oxygenase StyA (44). The high level of efficiency of self-sufficient cytochrome P450 enzymes compared to that of multicomponent types is attributed to the closer location between the heme-containing P450 domain (oxygenase) and a reductase domain (FAD/flavin mononucleotide [FMN] and NADH binding site), which should also promote the efficiency of diffusive transfer (38). These self-sufficient P450 systems are of high biocatalytic interest (8, 34, 39), and it is likely that other types of self-sufficient monooxygenases (e.g., flavoenzymes) behave in a similar way.Members of the gram-positive genus Rhodococcus are characterized by their exceptionally high level of metabolic versatility toward a broad range of organic substrates (31). Large genome sizes, the presence of megaplasmids, and a distinct gene redundancy (multiple enzyme homologs) favor these organisms as a promising source for novel enzymes (59). Moreover, several studies have provided evidence of functionally convergent evolution of the catabolic activities of rhodococci and proteobacteria (12, 37). Since most research on bacterial catabolic activities has so far been performed on the latter group, novel enzymes and mechanisms are likely to be identified in gram-positive bacteria.The nocardioform actinobacterium Rhodococcus opacus 1CP was originally isolated from contaminated soil by enrichment with 4-chloro- and 2,4-dichlorophenol as the sole carbon sources (18). To solve questions related to the enzymes involved in chlorophenol mineralization, an R. opacus 1CP clone library was generated, leading to the identification of a 9-kb genomic fragment harboring genes with a presumed function in styrene metabolism. Sequence analysis indicated the presence of an open reading frame (ORF) encoding a fusion protein composed of an oxygenase and a reductase subunit with a high level of similarity to the corresponding subunits of two-component SMOs from pseudomonads. Recombinant expression and biochemical characterization confirmed that it has enantioselective styrene epoxidation ability and showed that StyA2B is the first representative of a new class of NADH- and flavin-dependent single-component flavoprotein monooxygenases.  相似文献   

6.
A novel nitrilase superfamily amidase gene, designated azl13, was cloned from Streptomyces sp. 211726. Bioinformatic and biochemical analysis indicated that Azl13 belongs to a new subfamily in branch 13 of the nitrilase superfamily. His6-Azl13 was expressed in Escherichia coli BL21(DE3) and had the expected molecular mass of 31 kDa, and the enzymatic activity was best at 40 °C, pH 8.0. His6-Azl13 had amidase, aryl acylamidase, and acyl transferase activities, and it displayed an unusually wide substrate spectrum. His6-Azl13 was most active on 4-guanidinobutyramide, which is probably its natural substrate, moderately active on short-chain aliphatic amides and weakly active hydrolyzing aromatic and heterocyclic amides. His6-Azl13 also catalyzed acyl transfer to hydroxylamine from acetamide or the herbicide propanil. The substrate spectrum differs from that of the Pseudomonas amidase RamA, probably reflecting high salinity adaptation. The broad substrate spectrum of Azl13 is potentially useful for chemical synthesis and biodegradation.  相似文献   

7.
Rahier A 《Steroids》2011,76(4):340-352
Sterols become functional only after removal of the two methyl groups at C-4. This review focuses on the sterol C-4 demethylation process in higher plants. An intriguing aspect in the removal of the two C-4 methyl groups of sterol precursors in plants is that it does not occur consecutively as it does in yeast and animals, but is interrupted by several enzymatic steps. Each C-4 demethylation step involves the sequential participation of three individual enzymatic reactions including a sterol methyl oxidase (SMO), a 3β-hydroxysteroid-dehydrogenase/C4-decarboxylase (3βHSD/D) and a 3-ketosteroid reductase (SR). The distant location of the two C-4 demethylations in the sterol pathway requires distinct SMOs with respective substrate specificity. Combination of genetic and molecular enzymological approaches allowed a thorough identification and functional characterization of two distinct families of SMOs genes and two 3βHSD/D genes. For the latter, these studies provided the first molecularly and functionally characterized HSDs from a short chain dehydrogenase/reductase family in plants, and the first data on 3-D molecular interactions of an enzyme of the postoxidosqualene cyclase sterol biosynthetic pathway with its substrate in animals, yeast and higher plants. Characterization of these three new components involved in C-4 demethylation participates to the completion of the molecular inventory of sterol synthesis in higher plants.  相似文献   

8.
Despite promising preclinical results, transient single-factor-based therapeutic angiogenesis has shown no definitive benefits in clinical trials. The use of skin-derived microorgans (SMOs), capable of sustained expression of angiogenic factors and sustained viability with their cellular and extracellular elements, constitutes an attractive alternative. We sought to evaluate the efficacy of SMO implantation in a porcine model of chronic myocardial ischemia. Eighteen pigs underwent placement of an ameroid constrictor on the proximal circumflex artery. Three weeks later, split-thickness skin biopsies were harvested and pigs were randomized to lateral wall implantation of either 8 or 16 SMOs or blank injections. The procedure was safe and resulted in no adverse events. Three weeks after treatment, SMO implantation resulted in significant improvement of lateral wall perfusion during pacing, assessed by isotope-labeled microspheres [post- vs. pretreatment ratios of lateral/anterior wall blood flow were 1.31 +/- 0.09 (SMOs) and 1.04 +/- 0.06 (controls); P = 0.03]. No significant difference in angiographic scores was observed. Microvascular relaxation in response to VEGF was impaired in the ischemic territory of the control group but returned to normal after SMO implantation, indicating restoration of endothelial function. Molecular studies showed significant increases in VEGF and CD31 expression in the ischemic area of treated animals. Morphometric analysis showed increased neovascularization with SMO treatment. Autotransplantation of SMOs constitutes a novel approach for safe and effective therapeutic angiogenesis with improvement in perfusion, normalization of microvascular reactivity, and increased expression of VEGF and CD31.  相似文献   

9.
ABSTRACT: BACKGROUND: Polyamine oxidase enzymes catalyze the oxidation of polyamines and acetylpolyamines. Since polyamines are basic regulators of cell growth and proliferation, their homeostasis is crucial for cell life. Members of the polyamine oxidase gene family have been identified in a wide variety of animals, including vertebrates, arthropodes, nematodes, placozoa, as well as in plants and fungi. Polyamine oxidases (PAOs) from yeast can oxidize spermine, N1-acetylspermine, and N1-acetylspermidine, however, in vertebrates two different enzymes, namely spermine oxidase (SMO) and acetylpolyamine oxidase (APAO), specifically catalyze the oxidation of spermine, and N1-acetylspermine/N1-acetylspermidine, respectively. Little is known about the molecular evolutionary history of these enzymes. However, since the yeast PAO is able to catalyze the oxidation of both acetylated and non acetylated polyamines, and in vertebrates these functions are addressed by two specialized polyamine oxidase subfamilies (APAO and SMO), it can be hypothesized an ancestral reference for the former enzyme from which the latter would have been derived. RESULTS: We analysed 36 SMO, 26 APAO and 14 PAO homologue protein sequences from 54 taxa including various vertebrates and invertebrates. The analysis of the full-length sequences and the principal domains of vertebrate and invertebrate PAOs yielded consensus primary protein sequences for vertebrate SMOs and APAOs, and invertebrate PAOs. This analysis, coupled to molecular modeling techniques, also unveiled sequence regions that confer specific structural and functional properties, including substrate specificity, by the different PAO subfamilies. Molecular phylogenetic trees revealed a basal position of all the invertebrates PAO enzymes relative to vertebrate SMOs and APAOs. PAOs from insects constitute a monophyletic clade. Two PAO variants sampled in the amphioxus are basal to the dichotomy between two well supported monophyletic clades including, respectively, all the SMOs and APAOs from vertebrates. The two vertebrate monophyletic clades clustered strictly mirroring the organismal phylogeny of fishes, amphibians, reptiles, birds, and mammals. Evidences from comparative genomic analysis, structural evolution and functional divergence in a phylogenetic framework across Metazoa suggested an evolutionary scenario where the ancestor PAO coding sequence, present in invertebrates as an orthologous gene, has been duplicated in the vertebrate branch to originate the paralogous SMO and APAO genes. A further genome evolution event concerns the SMO gene of placental, but not marsupial and monotremate, mammals which increased its functional variation following an alternative splicing (AS) mechanism. CONCLUSIONS: In this study the explicit integration in a phylogenomic framework of phylogenetic tree construction, structure prediction, and biochemical function data/prediction, allowed inferring the molecular evolutionary history of the PAO gene family and to disambiguate paralogous genes related by duplication event (SMO and APAO) and orthologous genes related by speciation events (PAOs, SMOs/APAOs). Further, while in vertebrates experimental data corroborate SMO and APAO molecular function predictions, in invertebrates the finding of a supported phylogenetic clusters of insect PAOs and the co-occurrence of two PAO variants in the amphioxus urgently claim the need for future structure-function studies.  相似文献   

10.
A recombinant Escherichia coli (pBAB1) containing styrene monooxygenase (SMO) was developed for the conversion of styrene to enantiopure (S)-styrene oxide that is an important chiral building block in organic synthesis. The styAB genes encoding SMO was cloned into a multicopy plasmid under the tightly regulated promoter of bacterial l-arabinose operon which is inducible by l-arabinose. The recombinant showed that expression level of StyA protein and whole-cell SMO activities were varied depending on the concentration of the inducer l-arabinose. The maximum SMO activity was 108 U/g cdw when the cells were induced with 0.2% l-arabinose. SDS-PAGE and Western blot analyses indicated that whole-cell SMO activity was strongly correlated with the expression level of StyA protein. Organic-aqueous two-phase experiment could yield 50 mM enantiopure (S)-styrene oxide in organic phase in 18 h, but the recombinant SMO activity was unstable during the reaction. The expression of styAB under the control of l-arabinose promoter was significantly repressed in the presence of glucose.  相似文献   

11.
Styrene can efficiently be oxidized to (S)-styrene oxide by recombinant Escherichia coli expressing the styrene monooxygenase genes styAB from Pseudomonas sp. strain VLB120. Targeting microbial physiology during whole-cell redox biocatalysis, we investigated the interdependency of styrene epoxidation, growth, and carbon metabolism on the basis of mass balances obtained from continuous two-liquid-phase cultures. Full induction of styAB expression led to growth inhibition, which could be attenuated by reducing expression levels. Operation at subtoxic substrate and product concentrations and variation of the epoxidation rate via the styrene feed concentration allowed a detailed analysis of carbon metabolism and bioconversion kinetics. Fine-tuned styAB expression and increasing specific epoxidation rates resulted in decreasing biomass yields, increasing specific rates for glucose uptake and the tricarboxylic acid (TCA) cycle, and finally saturation of the TCA cycle and acetate formation. Interestingly, the biocatalysis-related NAD(P)H consumption was 3.2 to 3.7 times higher than expected from the epoxidation stoichiometry. Possible reasons include uncoupling of styrene epoxidation and NADH oxidation and increased maintenance requirements during redox biocatalysis. At epoxidation rates of above 21 μmol per min per g cells (dry weight), the absence of limitations by O2 and styrene and stagnating NAD(P)H regeneration rates indicated that NADH availability limited styrene epoxidation. During glucose-limited growth, oxygenase catalysis might induce regulatory stress responses, which attenuate excessive glucose catabolism and thus limit NADH regeneration. Optimizing metabolic and/or regulatory networks for efficient redox biocatalysis instead of growth (yield) is likely to be the key for maintaining high oxygenase activities in recombinant E. coli.  相似文献   

12.
The microbial degradation of 10 linear unsaturated dimers (I to IV) prepared from styrene and o-, m-, or p-methylstyrene was investigated with two soil bacteria, Alcaligenes sp. strain 559 and Pseudomonas sp. strain 419. The two strains decomposed styrene dimer I and all styrene-methylstyrene codimers II and III, but methylstyrene homodimers IV remained intact. The degradation rates of codimers II and III of o- and m-methylstyrenes were found to depend on both their structure and the strain used; i.e., Alcaligenes sp. strain 559 decomposed III faster than II, whereas the reverse order (II > III) was obtained with Pseudomonas sp. strain 419. In biodegradation by the former strain, the codimers were degraded faster in the presence of styrene dimer I than in its absence, but no such effect of dimer I was observed with the latter.  相似文献   

13.
The gastric pathogen Helicobacter pylori causes peptic ulcer disease and gastric cancer. We have reported that in H. pylori-activated macrophages, nitric oxide (NO) derived from inducible NO synthase (iNOS) can kill the bacterium, iNOS protein expression is dependent on uptake of its substrate l-arginine (l-Arg), the polyamine spermine can inhibit iNOS translation by inhibiting l-Arg uptake, and inhibition of polyamine synthesis enhances NO-mediated bacterial killing. Because spermine oxidase (SMO), which back-converts spermine to spermidine, is induced in macrophages by H. pylori, we determined its role in iNOS-dependent host defense. SMO shRNA knockdown in RAW 264.7 murine macrophages resulted in a marked decrease in H. pylori-stimulated iNOS protein, but not mRNA expression, and a 90 % reduction in NO levels; NO production was also inhibited in primary murine peritoneal macrophages with SMO knockdown. There was an increase in spermine levels after H. pylori stimulation that rapidly decreased, while SMO knockdown caused a greater increase in spermine that was sustained. With SMO knockdown, l-Arg uptake and killing of H. pylori by macrophages was prevented. The overexpression of SMO by transfection of an expression plasmid prevented the H. pylori-stimulated increase in spermine levels, and led to increased l-Arg uptake, iNOS protein expression and NO production, and H. pylori killing. In two human monocytic cell lines, U937 and THP-1, overexpression of SMO caused a significant enhancement of NO production with H. pylori stimulation. By depleting spermine, SMO can abrogate the inhibitory effect of polyamines on innate immune responses to H. pylori by enhancing antimicrobial NO production.  相似文献   

14.
Nine different sulfur-containing compounds were biotransformed to the corresponding sulfoxides by Escherichia coli Bl21(DE3) cells expressing styrene monooxygenase (SMO) from Pseudomonas putida CA-3. Thioanisole was consumed at 83.3 μmoles min?1 g cell dry weight?1 resulting mainly in the formation of R-thioanisole sulfoxide with an enantiomeric excess (ee) value of 45 %. The rate of 2-methyl-, 2-chloro- and 2-bromo-thioanisole consumption was 2-fold lower than that of thioanisole. Surprisingly, the 2-methylthioanisole sulfoxide product had the opposite (S) configuration to that of the other 2-substituted thioanisole derivatives and had a higher ee value (84 %). The rate of oxidation of 4-substituted thioanisoles was higher than the corresponding 2-substituted substrates but the ee values of the products were consistently lower (10–23 %). The rate of benzo[b]thiophene and 2-methylbenzo[b]thiophene sulfoxidation was approximately 10-fold lower than that of thioanisole. The ee value of the benzo[b]thiophene sulfoxide could not be determined as the product racemized rapidly. E. coli cells expressing an engineered SMO (SMOeng R3-11) oxidised 2-substituted thioanisoles between 1.8- and 2.8-fold faster compared to cells expressing the wild-type enzyme. SMOeng R3-11 oxidised benzo[b]thiophene and 2-methylbenzo[b]thiophene 10.1 and 5.6 times faster that the wild-type enzyme. The stereospecificity of the reaction catalysed by SMOeng was unchanged from that of the wild type. Using the X-ray crystal structure of the P. putida S12 SMO, it was evident that the entrance of substrates into the SMO active site is limited by the binding pocket bottleneck formed by the side chains of Val-211 and Asn-46 carboxyamide group.  相似文献   

15.
Bacterial atrazine catabolism is initiated by the enzyme atrazine chlorohydrolase (AtzA) in Pseudomonas sp. strain ADP. Other triazine herbicides are metabolized by bacteria, but the enzymological basis of this is unclear. Here we begin to address this by investigating the catalytic activity of AtzA by using substrate analogs. Purified AtzA from Pseudomonas sp. strain ADP catalyzed the hydrolysis of an atrazine analog that was substituted at the chlorine substituent by fluorine. AtzA did not catalyze the hydrolysis of atrazine analogs containing the pseudohalide azido, methoxy, and cyano groups or thiomethyl and amino groups. Atrazine analogs with a chlorine substituent at carbon 2 and N-alkyl groups, ranging in size from methyl to t-butyl, all underwent dechlorination by AtzA. AtzA catalyzed hydrolytic dechlorination when one nitrogen substituent was alkylated and the other was a free amino group. However, when both amino groups were unalkylated, no reaction occurred. Cell extracts were prepared from five strains capable of atrazine dechlorination and known to contain atzA or closely homologous gene sequences: Pseudomonas sp. strain ADP, Rhizobium strain PATR, Alcaligenes strain SG1, Agrobacterium radiobacter J14a, and Ralstonia picketti D. All showed identical substrate specificity to purified AtzA from Pseudomonas sp. strain ADP. Cell extracts from Clavibacter michiganensis ATZ1, which also contains a gene homologous to atzA, were able to transform atrazine analogs containing pseudohalide and thiomethyl groups, in addition to the substrates used by AtzA from Pseudomonas sp. strain ADP. This suggests that either (i) another enzyme(s) is present which confers the broader substrate range or (ii) the AtzA itself has a broader substrate range.  相似文献   

16.
Pseudomonas sp. VLB120 was isolated in Stuttgart, Germany, as a styrene degrading organism. The complete genome sequence includes genomic information of solvent tolerance mechanisms, metabolic pathways for various organic compounds, and the megaplasmid pSTY.  相似文献   

17.
We developed a biocatalyst by cloning the styrene monooxygenase genes (styA and styB) from Pseudomonas fluorescens ST responsible for the oxidation of styrene to its corresponding epoxide. Recombinant Escherichia coli was able to oxidize different aryl vinyl and aryl ethenyl compounds to their corresponding optically pure epoxides. The results of bioconversions indicate the broad substrate preference of styrene monooxygenase and its potential for the production of several fine chemicals.  相似文献   

18.
19.
Deuterated styrene ([2H8]styrene) was used as a tracer in combination with phospholipid fatty acid (PLFA) analysis for characterization of styrene-degrading microbial populations of biofilters used for treatment of waste gases. Deuterated fatty acids were detected and quantified by gas chromatography-mass spectrometry. The method was evaluated with pure cultures of styrene-degrading bacteria and defined mixed cultures of styrene degraders and non-styrene-degrading organisms. Incubation of styrene degraders for 3 days with [2H8]styrene led to fatty acids consisting of up to 90% deuterated molecules. Mixed-culture experiments showed that specific labeling of styrene-degrading strains and only weak labeling of fatty acids of non-styrene-degrading organisms occurred after incubation with [2H8]styrene for up to 7 days. Analysis of actively degrading filter material from an experimental biofilter and a full-scale biofilter by this method showed that there were differences in the patterns of labeled fatty acids. For the experimental biofilter the fatty acids with largest amounts of labeled molecules were palmitic acid (16:0), 9,10-methylenehexadecanoic acid (17:0 cyclo9-10), and vaccenic acid (18:1 cis11). These lipid markers indicated that styrene was degraded by organisms with a Pseudomonas-like fatty acid profile. In contrast, the most intensively labeled fatty acids of the full-scale biofilter sample were palmitic acid and cis-11-hexadecenoic acid (16:1 cis11), indicating that an unknown styrene-degrading taxon was present. Iso-, anteiso-, and 10-methyl-branched fatty acids showed no or weak labeling. Therefore, we found no indication that styrene was degraded by organisms with methyl-branched fatty fatty acids, such as Xanthomonas, Bacillus, Streptomyces, or Gordonia spp.  相似文献   

20.
The use of bioluminescence as a sensitive marker for detection of Pseudomonas spp. in the rhizosphere was investigated. Continuous expression of the luxCDABE genes, required for bioluminescence, was not detectable in the rhizosphere. However, when either a naphthalene-inducible luxCDABE construct or a constitutive luxAB construct (coding only for the luciferase) was introduced into the Pseudomonas cells, light emission could be initiated just prior to measurement by the addition of naphthalene or the substrate for luciferase, n-decyl aldehyde, respectively. These Pseudomonas cells could successfully be detected in the rhizosphere by using autophotography or optical fiber light measurement techniques. Detection required the presence of 103 to 104 CFU/cm of root, showing that the bioluminescence technique is at least 1,000-fold more sensitive than β-galactosidase-based systems.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号